bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Defining the RBPome of T helper cells to study higher order post-transcriptional

regulation

Kai P. Hoefig1*, Alexander Reim2*, Christian Gallus3*, Elaine H. Wong4, Gesine Behrens1,

Christine Conrad4, Meng Xu1, Taku Ito-Kureha4, Kyra Defourny4,10, Arie Geerlof5, Josef

Mautner6, Stefanie M. Hauck7, Dirk Baumjohann4,11, Regina Feederle8, Matthias Mann2,

Michael Wierer2,12#, Elke Glasmacher3,9#, Vigo Heissmeyer1,4#

1 Research Unit Molecular Immune Regulation, Helmholtz Center Munich, Munich, Germany

2 Department of Proteomics and Signal Transduction, Max-Planck-Institute of Biochemistry,

Munich, Germany

3 Institute of Diabetes and Obesity, Helmholtz Center Munich, Munich, Germany

4 Institute for Immunology at the Biomedical Center, Ludwig-Maximilians-Universität

München, Planegg-Martinsried, Germany

5 Institute of Structural Biology, Helmholtz Center Munich, Neuherberg, Germany

6 Research Unit Gene Vectors, Helmholtz Center Munich & Children’s Hospital, TU Munich,

Germany

7 Research Unit Science, Helmholtz Center Munich, Munich, Germany

8 Monoclonal Core Facility and Research Group, Institute for Diabetes and

Obesity, Helmholtz Center Munich, Neuherberg, Germany.

9 present address: Roche Pharma Research and Early Development, Large Molecule

Research, Roche Innovation Center Munich, Penzberg, Germany

10 present address: Department of Biomolecular Health Sciences, Utrecht University,

Utrecht, the Netherlands

11 present address: Medical Clinic III for Oncology, Immuno-Oncology and Rheumatology

University Hospital Bonn, University of Bonn, Bonn, Germany

1 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

12 present address: Proteomics Research Infrastructure, University of Copenhagen,

Copenhagen, Denmark

* These authors contributed equally

# Corresponding Authors: [email protected]

[email protected]

[email protected]

One sentence statement:

We provide an atlas of RNA-binding in human and mouse T helper cells as a

resource for studying higher order post-transcriptional gene regulation.

Abstract

Post-transcriptional gene regulation is complex, dynamic and ensures proper T cell function.

The targeted transcripts can simultaneously respond to various factors as evident for Icos,

an mRNA regulated by several RNA binding proteins (RBPs), including Roquin. However,

fundamental information about the entire RBPome involved in post-transcriptional gene

regulation in T cells is lacking. Here, we applied global RNA interactome capture (RNA-IC)

and orthogonal organic phase separation (OOPS) to human and mouse primary T cells and

identified the core T cell RBPome. This defined 798 mouse and 801 human proteins as

RBPs, unexpectedly containing signaling proteins like Stat1, Stat4 and Vav1. Based on the

vicinity to Roquin-1 in proximity labeling experiments, we selected ~50 RBPs for testing

coregulation of Roquin targets. Induced expression of these candidate RBPs in wildtype and

Roquin-deficient T cells unraveled several Roquin-independent contributions, but also

revealed Celf1 as a new Roquin-1-dependent and target-specific coregulator of Icos.

2 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

T lymphocytes as central entities of the adaptive immune system must be able to make

critical cell fate decisions fast. To exit quiescence, commit to proliferation and differentiation,

exert effector functions or form memory they strongly depend on programs of gene

regulation. Accordingly, they employ extensive post-transcriptional regulation through RBPs

and miRNAs or 3’ end oligo-uridylation and m6A RNA modifications. These RBPs, or RBPs

that recognize the modifications, directly affect the expression of by controlling mRNA

stability or translation efficiency 1, 2, 3, 4, 5. The studies in T helper cells have focused on a

small number of RNA-binding proteins, including HuR and TTP/Zfp36l1/Zfp36l2 6, 7, 8, 9 ,

Roquin-1/2 10, 11 and Regnase-1/4 12, 13 as well as some miRNAs like miR-17~92, miR-155,

miR-181, miR-125 or miR-146a 14. Moreover, first evidence for m6A RNA methylation in this

cell type has been provided 4. Underscoring the relevance for the immune system, loss-of-

function of these factors has often been associated with profound alterations in T cell

development or functions which caused immune-related diseases 14, 15, 16, 17. Intriguingly,

many key factors of the immune system have acquired long 3’-UTRs enabling their

regulation by multiple, and often overlapping sets of post-transcriptional regulators 18. Some

RBPs recruit additional co-factors as for example Roquin binding with Nufip2 to RNA 19 and

some of them have antagonistic RBPs like HuR and TTP 20 or Regnase-1 and Arid5a 21.

Such functional or physical interactions together with interdependent binding to the

transcriptome create enormous regulatory potential. The challenge is therefore to integrate

our current knowledge about individual RBPs into concepts of higher order gene regulation

that reflect the interplay of different, and ideally of all cellular RBPs.

A prerequisite for studying higher order post-transcriptional networks is to know cell type

specific RBPomes to account for differential expression and RBP plasticity. To this end

several global methods have been developed over the last decade, revealing a growing

number of RBPs that may even exceed recent estimates of ~7.5% of the human proteome

22. RNA interactome capture (RNA-IC) 23 is one widely used, unbiased technique, however, it

is constricted by design, exclusively identifying proteins binding to polyadenylated RNAs. In

3 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

contrast, orthogonal organic phase separation (OOPS) analyzes all UV-crosslinked protein-

RNA adducts from interphases after organic phase separation.

The interactions of RBPs with RNA typically involve charge-, sequence- or structure-

dependent interactions and to date over 600 structurally different RNA-binding domains

(RBD) have been identified in canonical RBPs of the human proteome 22. However, global

methods also identified hundreds of non-canonical RBPs, which oftentimes contained

intrinsically disordered regions (IDRs). Surprisingly, as many as 71 human proteins with well-

defined metabolic functions were found to interact with RNA 24 introducing the concept of

“moonlighting”. Depending on availability from their “day job” in metabolism such proteins

also bear the potential to impact on RNA regulation. Recent large-scale approaches have

increased the number of EuRBPDB-listed human RBPs to currently 2949 25 , suggesting that

numerous RNA/RBPs interactions and cell-type specific gene regulations have gone

unnoticed so far.

As a first step towards a global understanding of post-transcriptional gene regulation we

experimentally defined all proteins that can be crosslinked to RNA in T helper cells. RNA-IC

and OOPS identified 310 or 1200 proteins in primary CD4+ T cells interacting with

polyadenylated transcripts or all RNA species, respectively. Importantly, this dataset now

enables the study of higher order gene regulation by determining for example how the

cellular RBPs participate or intervene with post-transcriptional control of targets by individual

RBPs.

Results

Icos exhibits simultaneous and temporal regulation through several RBPs

A prominent example for complex post-transcriptional gene regulation is the mRNA encoding

inducible T-cell costimulator (Icos). It harbors a long 3’-UTR, which responds in a redundant

manner to Roquin-1 and Roquin-2 proteins, is repressed by Regnase-1 and microRNAs, but

also contains TTP binding sites 7, 10, 11, 13, 26, 27, 28. Moreover, the Icos 3’-UTR was proposed to

be modified by m6A methylation 29, which could either attract m6A-specific RBPs with YTH

4 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

domains 30, recruit or repel other RBPs 31, or interfere with base-pairing and secondary

structure- or miRNA/mRNA-duplex formation 32. Because of transcriptional and post-

transcriptional regulation, Icos expression exhibits a hundred-fold upregulation on the protein

level during T cell activation, which then declines after removal of the TCR stimulus (Fig. 1).

To determine the temporal impact on Icos expression by Roquin, Regnase-1, m6A and

miRNA regulation we analyzed inducible, CD4-specific inactivation of Roquin-1 together with

Roquin-2 (Rc3h1-2), Regnase-1 (Zc3h12a) or Wtap, an essential component of the m6A

methyltransferase complex 33, as well as Dgcr8, which is required for pre-miRNA biogenesis

34. To this end, we performed tamoxifen gavage on mice expressing a Cre-ERT2 knockin

allele in the CD4 locus 35 together with the floxed, Roquin paralogs encoding, Rc3h1 and

Rc3h2 alleles (Fig. 1a-c), Regnase-1 encoding Zc3h12a alleles (Fig. 1d-f) as well as Wtap,

(Fig. 1g-i) and Dgcr8 alleles, (Fig. 1j-l). We isolated CD4+ T cells from these mice and

expanded them for five days. Confirming target deletion on the protein level (Fig. 1c, f, i, l)

we determined a strong negative effect on Icos expression by Roquin and Regnase-1 on

days 2-5 (Fig. 1a-b and d-e), a moderate positive effect of Wtap on days 2-5 (Fig. 1g-h),

and only a small effect of Dgcr8, with an initial tendency of negative (day 1) and later on

positive effects (days 4-5) (Fig. 1j-k). We next asked whether T cell activation affects the

expression levels of known regulators of Icos, as well as other RBPs, to install temporal

compartmentalization. To do so, we monitored the expression of a panel of RBPs in mouse

CD4+ T cells over the same time course (Fig. 1m-q). Indeed, we observed fast upregulation

of RBPs as determined with pan-Roquin, Nufip2 19, Fmrp, Fxr1, Fxr2, pan-TTP/Zfp36 7, pan-

Ythdf (Supplementary Fig. 1a), or Celf1 specific , but also slower accumulation

as determined with Regnase-1 or Rbms1 specific antibodies (Fig. 1m, o). There was also

downregulation of RBPs as shown with pan-AGO 36 and Cpeb4 specific antibodies (Fig. 1m-

n and p). Of note, we also observed signs of post-translational regulation showing

incomplete or full cleavage of Roquin or Regnase-1 proteins 10, 13, respectively (Fig. 1m),

and the induction of a slower migrating band for Celf1, likely phosphorylation 37 (Fig. 1q).

Factors with the potential to cooperate, as involved for Roquin and Nufip219 or Roquin and 5 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Regnase-110, showed overlapping temporal regulation (Fig. 1m), suggesting reinforcing

effects on the Icos target. Together these data indicate that mRNA targets respond to

simultaneous inputs from several RBPs, which are aligned by dynamic expression and post-

translational regulation to orchestrate redundant, cooperative and antagonistic effects into a

higher order regulation.

The polyadenylated RNA-bound proteome of mouse and human T helper cells

To decipher this post-transcriptional network we set out to determine the RNA-binding

protein signature in primary mouse and human CD4+ T cells. To identify mRNA-binding

+ proteins we first performed mRNA capture experiments on CD4 T cells expanded under TH0

culture conditions (Fig. 2a). Pull-down with oligo-dT beads enabled the enrichment of

mRNA-bound proteins as determined by silver staining, which were increased in response to

preceding UV irradiation of the cells (Supplementary Fig. 1b). Reverse transcription

quantitative PCR (RT-qPCR) confirmed the recovery of specific mRNAs, such as for the

housekeeping genes Hprt and β-actin. Both mRNAs were enriched at least 2-3 fold after UV

crosslink, but there was no detection of non-polyadenylated 18S rRNA (Supplementary Fig.

1c). Focusing on protein recovery, we determined greatly enriched polypyrimidine tract-

binding protein 1 (Ptbp1) RBP compared to the negative control β-tubulin in mRNA capture

experiments using the EL-4 thymoma cell line (Supplementary Fig. 1d). Next we performed

mass spectrometry (MS) on captured proteins from murine and human T cells (Fig. 2b-c).

Quantifying proteins bound to mRNA in crosslinked (CL) versus non-crosslinked (nCL)

samples we defined a total of 312 mouse (Fig. 2b) and 308 human mRNA-binding proteins

(mRBPs) (Fig. 2c) with an overlap of ~70% (Supplementary Table 1), which is in

concordance with the overlap of all listed RBPs for these two species in the eukaryotic RBP

database (http://EuRBPDB.syshospital.org). (GO) analysis identified the term

‘mRNA binding’ as most significantly enriched (Fig. 2d). The top ten GO terms in mouse

were also strongly enriched in the human dataset with comparable numbers of proteins

assigned to the individual GO terms in both species (Fig. 2d). RBPs not only bind RNA by

6 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

classical RBDs, but RNA-interactions can also map to intrinsically disordered regions (IDRs)

38. Furthermore, low complexity regions (LCRs) have also been reported to be

overrepresented in RBPs 23. Indeed, IDRs (Fig. 2e) and LCRs (Supplementary Fig. 1e)

were strongly enriched protein characteristics of mouse and human RNA-IC-identified

RBPomes. We next wondered how much variation exists in the composition of RBPs

between different T helper cell subsets. Repeating RNA-IC experiments in mouse and

human iTreg cells (Supplementary Fig. 2a), we find an overlap of 96% or 90% with the

respective mouse or human iTreg with Teff RBPomes, suggesting that the same RBPs bind

to the transcriptome in different T helper cell subsets (Table 1). Nevertheless, 47 or 48

proteins were exclusively identified in mouse or human effector T cells, respectively, and 10

or 28 proteins were only found in mouse or human Treg cells, respectively (Supplementary

Fig. 2b). Although these differences could indicate the existence of subset-specific RBPs,

they could also be related to an incomplete assessment of RBPs by the RNA-IC technology.

The global RNA-bound proteome of mouse and human T helper cells

We therefore employed a second RNA-centric method of orthogonal organic phase

separation (OOPS) to extend our definition of the RBPome and cross-validate our results.

Similar to interactome capture, OOPS preserves cellular protein/RNA interactions by UV

crosslinking of intact cells. The physicochemical properties of the resulting aducts direct

them towards the interphase in the organic and aqueous phase partitioning procedure (Fig.

3a). Following several cycles of interphase transfer and phase partitioning, RNase treatment

releases RNA-bound proteins into the organic phase, making them amenable to mass

spectrometry 39, 40. Evaluating the method, we investigated RNA and proteins from purified

interphases derived from CL and nCL MEF cell samples. RNAs with crosslinked proteins

purified from interphases hardly migrated into agarose gels, but regained normal migration

behavior after protease digest, as judged from the typical 18S and 28S rRNA pattern

(Supplementary Fig. 3a). Conversely, known RBPs like Roquin-1 and Gapdh appeared

after crosslinking in the interphase and could be recovered after RNase treatment from the

7 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

organic phase (Supplementary Fig. 3b). Utilizing the same cell numbers and culture

conditions of T cells, this method identified in total 1255 and 1159 significantly enriched

RBPs for mouse or human T cells, respectively, when comparing CL and nCL samples

(Supplementary Table 1). The overlap between both organisms was 55% (Fig. 3b) and

60% (Fig. 3c) in relation to the individual mouse and human RBPomes. Although

glycosylated proteins are known to also accumulate in the interphase 39, we experimentally

verified that they did not migrate into the organic phase after RNase treatment

(Supplementary Fig. 3c-d). Analyzing OOPS-derived RBPomes for gene ontology

enrichment using the same approach as for RNA-IC the GO term ‘mRNA binding’ was again

most significantly enriched in mouse and man (Fig. 3d). The top 10 GO terms were RNA

related and six of them overlapped with those identified for RNA-IC-derived RBPomes. High

similarity between mouse and human RBPomes becomes apparent by the similarity in all

GO categories, including ‘molecular function’ (Fig. 3d), ‘biological process’ and ‘cellular

component’ (Supplementary Fig. 4). Although our OOPS approach exceeded by far the

quantity of RNA-IC identified RBPs, the number of ~1200 RBPs well matched published

RBPomes of HEK293 (1410 RBPs), U2OS (1267 RBPs) and MCF10A (1165 RBPs) cell

lines 39.

Defining the core T helper cell RBPome

To define a T helper cell RBPome we first made sure that RNA-IC and OOPS did not

preferentially identify high abundance proteins (Fig. 4a-b). In comparison to single shot total

proteomes OOPS-identified RBPs spanned the whole range of protein expression without

apparent bias. In general, this was also true for RNA-IC, with a slight tendency to more

abundantly expressed proteins. This however might be a true effect since messenger RNA

binding RBPs have been reported to be higher in expression even in comparison to other

RBPs 22. We used the recently established comprehensive eukaryotic RBP database as

reference to compare OOPS and RNA-IC-identified canonical and non-canonical RBPs from

mouse and human CD4+ T cells. The numbers of proteins in the mouse T helper cell

8 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

RBPomes created by RNA-IC and OOPS ranging from 312 to 1255 made up 10% to 40% of

all listed EuRBPDB proteins, respectively (Fig. 4c). OOPS-identified T cell RBPs

outnumbered those from RNA-IC experiments by a factor of four, which was predominantly

due to the eight times higher number of non-canonical RBPs. Interestingly though, there

were also twice as many canonical RBPs significantly enriched by OOPS (Fig. 4c).

Analyzing the ten most abundantly annotated mouse RBDs (comprising 26 to 224 members)

showed that RNA-IC and OOPS often identified the same canonical RBPs (Supplementary

Fig. 5), however at least equal, higher or much higher numbers were detected in OOPS

samples depending on the specific RBD (Fig. 4d). These findings suggested that the OOPS

method recovered RBPs from additional, non-polyadenylated RNAs and that RNA-IC-

derived RBPomes are specific but incomplete. These conclusions are also supported by

highly similar results obtained for the human CD4+ T cell RBPome (Fig. 4e-f). In a 4-way

comparison of mouse and human RBPs identified by OOPS and RNA-IC (Fig. 4g) we

conservatively defined all proteins that were identified by at least two datasets as ‘core CD4+

T cell RBPomes’ discovering 798 mouse and 801 human RBPs in this category

(Supplementary Table 1). A sizable number of 519 mouse and 424 human proteins were

exclusively enriched by the OOPS method. While we cannot rule out false-positives, more

than 55% of the proteins of both subsets matched to EuRBPDB-listed annotations (not

shown). These findings suggested that genuine RBPs are found even outside of the

intersecting set of OOPs and RNA-IC identified proteins and that the definition of RBPomes

profits from employing different biochemical approaches.

T cell signaling proteins with unexpected RNA-binding function

Some of the identified RBPs of the core proteome including Stat1, Stat4 and Crip1 are not

expected to be associated with mRNA in cells. We therefore established an assay to confirm

RNA-binding of these candidates. To do so, GFP-tagged candidate proteins were

overexpressed in HEK293T cells, which were UV crosslinked, and extracts were used for

immunoprecipitations with GFP specific antibodies. By Western or North-Western blot

9 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

analyses with oligo(dT) probes we could verify pull-down of GFP-tagged proteins and the

association with mRNA for the RBPs Roquin-1 and Rbms1, but also for the lactate

dehydrogenase (Ldha) protein, a metabolic enzyme with known ability to also bind RNA 23

(Fig. 5a). Via this approach, the determined RNA association of Stat1, Stat4 and Crip1 by

global methods was indeed confirmed (Fig. 5b), but appeared less pronounced as

compared to prototypic RPBs, and similar with regard to Ldha (Fig 5a). This finding supports

a potential moonlighting function of these signaling proteins. As the identity of the associated

RNA species for these RBPs is unknown, we established a dual luciferase assay to

determine the impact of the different proteins on the expression of the renilla luciferase

transcript when they were tethered to an artificial 3’-UTR (Fig. 5c). We utilized the

λN/5xboxB system 41 and confirmed the expression of fusion proteins with a newly

established λN-specific antibody (Fig. 5d-e). Importantly, Stat1 and Stat4 repressed

luciferase function almost to the same extend as the known negative regulators Pat1b and

Roquin-1, or other known RBPs, such as Celf1, Rbms1 and Cpeb4 (Fig. 5f). λN-Crip1 and

λN-Vav1 expression did not exert a positive or negative influence, since their relative

luciferase expression appeared unchanged compared to cells transfected to only express

the λN polypeptide (Fig. 5f). These data suggest that the signaling proteins Stat1 and Stat4

that we defined here as part of the T helper cell RBPome not only have the capacity to bind

mRNA but can also exert RNA regulatory functions.

Analyzing higher order post-transcriptional regulation

We next devised an experimental strategy to uncover RBPs with the potential to antagonize

or cooperate with the Roquin-1 RBP in the repression of its target mRNAs by performing

‘BioID’ experiments (Fig. 6a). In this protein-centric, proximity-based labelling method we

expressed a Roquin-1 BirA* fusion protein to identify proteins that reside within a short

distance of approximately ~10 nm 42 in T cells (Fig. 6a). In this dataset we sought for

matches with the T cell RBPome (Fig. 4g) to identify proteins that shared the features,

‘RNA-binding’ and ‘Roquin-1 proximity’. We first verified that the mutated version of the biotin

10 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

ligase derived from E. coli (BirA*) which was N-terminally fused to Roquin-1 or GFP was

able to biotinylate residues in Roquin or other cellular proteins (Fig. 6b) but did not interfere

with the ability of Roquin-1 to downregulate Icos (Fig. 6c). Doxycycline-induced BirA*-

Roquin-1 compared to BirA*-GFP expression in CD4+ T cells significantly enriched biotin

labelling of 64 proteins (Supplementary Table 2), including Roquin-1 (Rc3h1) itself or

Roquin-2 (Rc3h2) (Fig. 6d) as well as previously identified Roquin-1 interactors and

downstream effectors, such as Ddx6 and Edc4 26, components of the Ccr4/Not complex 43, 44

and Nufip2 19 (Fig. 6d). More than half of all proteins in proximity to Roquin-1 were also part

of the defined RBPome (Fig. 6e-f). Increasing this BioID list with additional proteins that we

determined in proximity to Roquin-1 when establishing and validating the BioID method in

fibroblasts (Supplementary Fig. 6a-d), we arrived at 143 proteins (Supplementary Table

2) of which 96 (67%) were part of the RBPome (Supplementary Fig. 6e and

Supplementary Fig. 7a). Cloning of 46 candidate genes in the context of N-terminal GFP

fusions (Supplementary Fig. 8a) and generating retroviruses to overexpress candidate

proteins we analyzed their effect on endogenous Roquin-1 targets (Supplementary Fig.

8b). CD4+ T cells were used from mice with Rc3h1fl/fl;Rc3h2fl/fl;rtTA alleles in combination

with (iDKO) or without the CD4Cre-ERT2 allele (WT) allowing induced inactivation of

Roquin-1 and -2 by 4’-OH-tamoxifen treatment. Reflecting deletion, the Roquin-1 targets

Icos, Ox40, Ctla4, IκBNS and Regnase-1 became strongly derepressed in induced double-

knockout (iDKO) T cells (Supplementary Fig. 8c). This elevated expression was corrected

to wildtype levels in iDKO T cells that were retrovirally transduced and doxycycline-induced

to express GFP-Roquin-1 (Supplementary Fig. 8c). The target expression in WT T cells

was only moderately reduced through overexpression of GFP-Roquin-1 (Supplementary

Fig. 8c). For the majority of the 46 candidate genes induced expression in WT and iDKO

CD4+ T cells did not alter expression of the five analyzed Roquin-1 targets, exemplified here

by the results obtained for Vav1 (Fig. 7a-b). Interestingly, we identified a new function for

Rbms1 (transcript variant 2), specifically upregulating Ctla-4 (Fig. 7c-d). Furthermore, we

demonstrated that Cpeb4 strongly upregulates Ox40 and, most strikingly, in the same cells 11 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Cpeb4 repressed Ctla-4 levels (Fig. 7e-f). While these findings are new and noteworthy,

they occurred in a Roquin-1 independent manner. In contrast to these effects, we

determined a higher order regulation of Icos by Celf1. Induced Celf1 expression in wildtype T

cells clearly upregulated Icos, Ctla-4 and Ox40 expression and this function was obliterated

in Roquin-1-deficient iDKO cells (Fig. 7g-h). Importantly, this effect could not be explained

by Celf1-mediated repression of Roquin-1 on the protein or mRNA level (Supplementary

Fig. 8d-e). Together these findings demonstrate responsiveness of Roquin targets to many

other RBPs and a higher order, Roquin-dependent regulation of transcripts encoding for the

costimulatory receptors Icos, Ox40 and Ctla-4 by Celf1.

Discussion

The work on post-transcriptional gene regulation in T helper cells has focused on some

miRNAs and several RNA-binding proteins, and few reports described m6A RNA

methylation in this cell type. Although arriving at a more or less detailed understanding of

individual molecular relationships and regulatory circuits, this isolated knowledge assembles

into a very incomplete picture. Creating an atlas of the human and mouse T helper cell

RBPomes has now opened the stage, allowing to work towards understanding connections,

complexity and principles of post-transcriptional regulatory networks in these cells.

RNA-IC and OOPS are two complementary methods to define RBPs on a global scale.

While RNA-IC specifically queries for proteins bound to polyadenylated RNAs, OOPS

captures the RNA bound proteome in its whole. Applying both methods to T helper cells of

two different organisms allowed us to cross-validate the results from both methods and

solidified our description of the core mouse and human T helper cell RBPome. While the

vast majority of RNA-IC-identified CD4+ T cell RBPs were previously known RNA binders,

OOPS typically repeated and profoundly expanded these results (Supplementary Fig. 5),

with the possible caveat of identifying false-positive proteins. Contradicting this possibility at

least in part, more than half of OOPS identified proteins that were exclusively found in

mouse or man were EuRBPDB-listed.

12 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Strikingly, the signaling proteins Stat1 and Stat4 were identified by mouse RNA-IC and

human OOPS and were just below the cut-off in the mouse OOPS dataset, and we could

support their RBP function by additional RNA-binding assays. Undoubtedly, the defined

human and mouse T cell RBPomes contain many more unusual RBPs that would warrant

further investigations. We assume that even interactions of proteins without prototypic RBDs,

like the Stat proteins, with RNA will have consequences for both binding partners. As the

identity of the interacting mRNA(s) is unknown, we could only speculate about the post-

transcriptional impact. Nevertheless, Stat1 and Stat4 showed regulatory capacity in our

tethering assays. Intriguingly, RNA binding may also impact the function of Stat proteins as

transcription factors. Supporting this notion, early results found Stat1 bound to the non-

coding, polyadenylated RNA ‘TSU’, derived from a trophoblast cDNA library, and

translocation of Stat1 into the nucleus was reduced after TSU RNA microinjection into HeLa

cells 45, 46 .

Many 3’-UTRs, which effectively instruct post-transcriptional control, exhibit little sequence

conservation between species, and the exact modules which determine regulation are not

known. This for example is true for the Icos mRNA 19, 26. On the side of the trans-acting

factors, we find high similarity between the RBPomes of T helper cells of mouse and human

origin, actually reflecting the general overlap of so far determined RBPomes from many cell

lines of these species. We interpret this evolutionary conservation as holding ready similar

sets of RBPs in T helper cells across species, which are then able to work together on

composite cis-elements to reach comparable regulation of 3’-UTRs in the different

organisms, despite sequence variability and differences in the composition of elements.

We not only define the first RBPomes of human and mouse T helper cells. We also provide

potential avenues of how to make use of this information. Screening a set of candidates from

the T cell RBPome for effects on Roquin targets, our findings support a concept in which

post-transcriptional targets are separated into “regulons” 47. These regulons comprise

coregulated mRNA subsets responding to the same inputs and often functioning in the same

13 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

biological process. Therefore, complex and differential binding of targets by RBPs of the

cellular RBPome specifies the possible regulons and differential functions of the cell. Roquin

cooperated, coregulated, or antagonized in the regulation of Icos with Regnase-1, m6A

methylation and miRNA functions. Moreover, our screening approach added Rbms1 as

Roquin-independent and Celf1 as Roquin-dependent coregulator in the Icos containing

regulon. It also revealed that different targets responded very differently to the expression of

specific coregulators, as for example Ctla4 and Ox40 were inhibited or induced by the same

RBP, Cpeb4, respectively. Together these data indicated an unexpected wealth of possible

inputs from the T helper cell RBPome and suggested highly variable and combinatorial

mRNP compositions in higher order post-transcriptional gene regulation of individual targets.

To solve the seemingly simple question which RBPs regulate which mRNA in T helper cells,

we will require further knowledge about individual contributions, binding sites and composite

cis-elements, temporal and interdependent occupancies, interactions among RBPs and with

downstream effector molecules. In this endeavor global protein and RNA centric approaches

make fundamental contributions.

Acknowledgements

We thank Hemalatha Mutiah (Ludwig-Maximilians-Universität Munich) for screening

hybridoma supernatants and also Claudia Keplinger (Helmholtz Center Munich) for excellent

technical support. For the provision of mouse lines we would like to thank Marc Schmidt-

Supprian (Rc3h1), Wolfgang Wurst (Rc3h2), Robert Blelloch (Dgcr8flox), Mingui Fu

(Zc3h12aflox) and Thorsten Buch (Cd4-Cre-ERT2). The work was supported by the German

Research Foundation grants SPP-1935 (to VH), SFB-1054 projects A03 and Z02 as well as

HE3359/7-1 and HE3359/8-1 to V.H. as well as grants from the Wilhelm Sander, Fritz

Thyssen, Else Kroner-Fresenius and Deusche Krebshilfe foundations to V.H.. D.B. was

supported by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation)

14 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

under Emmy Noether Programme BA 5132/1-1 and BA 5132/1‐2 (252623821), as well as

SFB 1054 project B12 (210592381). D.B. is a member of Germany's Excellence Strategy

EXC2151 577 (390873048).

Author contribution

EG and VH conceived the idea for the project and together with MW and MM supervised the

experimental work. KPH and VH wrote the manuscript with contributions from EG, MW and

AR. KPH performed OOPS, BioID and T cell transductions with help from GB, KD, SMH, JM

and CC. CG conducted RNA-IC experiments and RNA-binding assays. AR and SMH

performed mass spectrometry and AR analyzed RBPome data. SMH and KPH analyzed the

BioID data. MX contributed the tethering assays and EW analyzed Icos regulation for which

AG, RF, TI-K and DB provided unpublished reagents.

Data availability statement

The data sets generated and/or analyzed during the current study are available from the

corresponding authors on reasonable request.

Methods

Isolation, in vitro cultivation and transduction of primary CD4+ T cells

For in vitro cultivation of primary murine CD4+ T cells, mice were sacrificed and spleen as

well as cervical, axillary, brachial, inguinal and mesenteric lymph nodes were dissected and

pooled. Organs were squished and passed through a 100 μm filter under rinsing with T cell

isolation buffer (PBS supplemented with 2% FCS and 1 mM EDTA). Erythrocytes were

eliminated by incubating cells with TAC-lysis buffer (13 mM Tris, 140 mM NH4Cl, pH 7.2) for

5 min at room temperature. CD4+ T cells were isolated by negative selection using

EasySep™ Mouse CD4+ T cell isolation Kit (STEMCELL) according to manufacturer’s

15 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

protocol. Purified CD4+ T cells were cultured in DMEM (Invitrogen) T cell culture medium

supplemented with 10% FCS (Gibco), 1% Pen-Strep (Thermo Fisher), 1% HEPES-Buffer

(Invitrogen), 1% non-essential amino acids (NEAA; Invitrogen) and 1 mM β-mercaptoethanol

(Invitrogen). For activation and differentiation under TH1 conditions the T cells were

stimulated with α-CD3 (0.5 µg/mL; cl. 2C11, in house production), α-CD28 (2.5 µg/mL; cl.

37.5N, in house production), 10 μg/mL α-IL-4 (cl. 11B11, in house production) and 10 ng/mL

IL-12 (BD Pharmingen) and cultured on goat α-hamster IgG (MP Biochemicals) pre-coated

6-or 12-well culture plates for 40-48h at an initial cell density of 5 or 1.5×106 cells/mL,

respectively. The cells were then resuspended and cultured in medium supplemented with

200 IE/mL recombinant hIL-2 (Novartis) in a 10% CO2 incubator and expanded for 2-4 days,

as indicated. Subsequently cells were fed with fresh IL-2-containing medium every 24h and

cultured at a density of 0.5-1×106 cells/mL. For in vitro deletion of floxed alleles of Rc3h1fl/fl

;Rc3h2fl/fl ;CD4Cre-ERT2;rtTA3 (but with no effect on the Cre-deficient WT control Rc3h1fl/fl

;Rc3h2fl/fl ;rtTA3) CD4+ T cells were treated with 1 µM 4´OH-tamoxifen (Sigma) for 24h prior

to activation at a cell density of 0,5-1×106 cells/mL. We performed retroviral transduction 40h

after the start of anti-CD3/CD28 activation, T cells were transduced with retroviral particles

using spin-inoculation (1h, 18 °C, 850 g), and after 4-6h co-incubation of T cells and virus,

virus particles were removed and T cells resuspended in T cell medium supplemented with

IL-2 as described before. To induce expression of pRetro-Xtight-GFP constructs in rtTA

expressing T cells, the transduced cells were cultured for 16h in the presence of doxycycline

(1 µg/mL) prior to flow cytometry analysis of expression of targets in transduced GFP+ cells.

In vivo deletion of loxP-flanked alleles and in vitro culture of CD4+ T cells

For in vitro culture analysis, deletion of Roquin (Rc3h1fl/fl;Rc3h2fl/fl;CD4-Cre-ERT2) 48,

Regnase-1 (Zc3h12afl/fl;CD4-Cre-ERT2) 49, Wtap (Wtapfl/fl;CD4-Cre-ERT2) 33, and Dgcr8

(Dgcr8fl/fl;CD4-Cre-ERT2) 34 encoding alleles in Cd4-cre-ERT2 mice was induced in vivo by

oral transfer of 5 mg tamoxifen (Sigma) in corn oil. Two doses of tamoxifen each day were

given on two consecutive days (total of 20 mg tamoxifen per mouse). Mice with the genotype

16 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

CD4-Cre-ERT2 (without floxed alleles) were used for wild-type controls. Mice were scarified

three days after the last gavage and total CD4+ T cells were isolated using the EasySepTM

Mouse T Cell Isolation Kit (Stem Cell) and activated under TH1 conditions as described

above.

Flow cytometry

Following in vivo deletion and CD4+ T cell isolation (above) cells were activated and cultured

under TH1 conditions. Cells were obtained daily for FACS analysis. The single cell

suspensions were stained with fixable violet dead cell stain (Thermo Fisher) for 20 min at

4°C. For the detection of surface proteins, cells were stained with the appropriate antibodies

in FACS buffer for 20 min at 4°C. After staining, cells were acquired on a FACS Canto II (3-

laser). The data were further processed with the software FlowJo 10 software (BD

Bioscience). The following antibodies were used: anti-CD4 (cl. GK1.5), anti-CD44 (cl. IM7),

anti-CD62L (cl. MEL-14), anti-Icos (cl. C398.4A), anti-Ox40 (cl. OX-86), all from eBioscience,

anti-CD25 (cl. PC61, Biolegend).

Effects of doxycycline-induced expression of 46 GFP-GOI fusion protein in 2x106 wild-type

and Roquin iDKO CD4+ T cells were analyzed on day 6 after isolation (compare Suppl. Fig.

8b). First, proteins were treated with a fixable blue dead cell stain (Invitrogen) and after

washing, stained in three panels to interrogate the surface expression of Icos and Ox40

(Icos-PE, clone 7E.17G9/Ox40-APC, clone OX-86; both eBioscience) and to intracellularly

measure Ctla4, IκBNS (Ctla4-PE, UC10-4B9; eBioscience/cl. 4C1 rat monoclonal; in house

production) as well as Regnase-1 (cl. 15D11 rat monoclonal; in house production). For

intracellular staining cells were fixed in 2% formaldehyde for 15 min at RT, permeabilized by

washing in Saponin buffer and stained with the appropriate antibodies for 1h at 4 °C. After

washing, anti-rat-AF647 antibody (cl. poly4054; Biolegend) was added for 30 min. After

additional rounds of Saponin- and FACS buffer washing, acquisition was performed using a

LSR Fortessa.

17 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Isolation and differentiation of effector and regulatory T cells for RNA-IC

Naïve CD4+ T cells were isolated by using Dyna- and Detachabeads (11445D and 12406D,

Invitrogen) from spleens and mesenteric lymph nodes of 8-12 week old C57BL/6J mice. For

iTreg culture, cells were additionally selected for CD62L+ with anti-CD62L (clone: Mel14)

coated beads. All cells were then activated with plate-bound anti-CD3 (using first anti-

hamster, 55397, Novartis, then anti-CD3 in solution: clone: 2C11H: 0.1 µg/mL) and soluble

anti-CD28 (clone: 37N: 1 µg/mL) and cultured in RPMI medium (supplemented with 10%

(vol/vol) FCS, β-mercaptoethanol (0.05 mM, Gibco), penicillin-streptomycin (100 U/ml,

Gibco), Sodium Pyruvate (1 mM, Lonza), Non-Essential Amino Acids (1x, Gibco), MEM

Vitamin Solution (1x, Gibco), Glutamax (1x, Gibco) and HEPES pH 7.2 (10 mM, Gibco)). For

iTreg cell differentiation we additionally added the following cytokines and blocking

antibodies: rmIL-2 and rmTGF-β (both: R&D Systems, 5 ng/ml), anti-IL-4 (clone: 11B11, 10

µg/ml) and anti-IFNγ (clone: Xmg-121, 10 µg/ml). All antibodies were obtained in

collaboration with and from Regina Feederle (Helmholtz Center Munich). After differentiation

for 36-48 h cells were expanded for 2-3 days. iTreg cells were cultured in RPMI and 2000

units Proleukin S (02238131, MP Biomedicals) and Teff cells with 200 U Proleukin S. We only

used iTreg cells for experiments if samples achieved at least 80% Foxp3 positive cells

(00552300, Foxp3 Staining Kit, BD Bioscience).

EL-4 T cells were cultured in the same medium as primary T cells. HEK293T cells were

cultured in DMEM (supplemented with 10% (vol/vol) FCS, penicillin-streptomycin (100 U/ml,

Gibco) and Hepes, pH 7.2 (10 mM, Gibco)).

RBPome capture

6 For RBPome capture for mass spectrometry, 20 x 10 Teff or iTreg cells were either lysed

directly (nonirradiated, control) in 1 ml lysis buffer from the µMACS mRNA Isolation Kit (130-

075-201, Miltenyi) or suspended in 1 ml PBS, dispensed on a 10 cm dish and UV irradiated

at 0.2 J/cm2 at 254 nm for 1 min, washed with PBS, pelleted and subsequently lysed (UV 18 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

irradiated) and mRNA was isolated from both samples with the µMACS mRNA Isolation Kit

according to the manufacturer’s instructions. RNAs and crosslinked proteins were eluted

6 with 70° C RNase-free H2O. For RBPome capture for Western blot analysis, 400 x 10 EL-4

T cells were either lysed directly in 8 ml lysis buffer (nonirradiated, control) or suspended in

16 ml PBS, dispensed on sixteen 10 cm dishes and UV irradiated at 0.2 J/cm2 at 254 nm for

1 min, washed with PBS, pelleted and subsequently lysed in 8 ml lysis buffer (UV irradiated)

and then mRNA was isolated from both samples with the µMACS mRNA Isolation Kit using

500 µl oligo(dT) beads per sample. Each sample was split and run over two M columns

(130-042-801, Miltenyi) and each column was eluted with two times 100 µl RNase-free H2O.

The eluate was concentrated in Amicon centrifugal filter units (UFC501008, Merck) to a final

volume of ∼25 µl. 8 µl Lämmli buffer (4x) with 10% (vol/vol) β-mercaptoethanol was added

and the sample boiled for 5 min at 95° C. For protein analysis, samples were either flash-

frozen in liquid nitrogen for MS analysis or Lämmli loading dye was added for subsequent

analysis by Western blotting or silver staining.

OOPS

20-30 x 106 mouse CD4+ T cells were isolated as described above and activated without

bias. Cells were washed in PBS once and resuspended in 1 ml of cold PBS and transferred

into one well of a six well dish. Floating on ice, the cells in the open 6 well plate were UV

irradiated once at 0.4 J/cm2 and twice at 0.2 J/cm2 at 254 nm with shaking in-between

sessions. After irradiation the 1ml of cell suspension was transferred into a FACS tube and

the well was washed with another 1 ml of cold PBS which was also added to the FACS tube.

After centrifugation the cell pellet was completely dissolved in 1 ml of Trizol (Sigmal). The

remainder of the procedure was performed strictly according to 39 with the exception that we

broke up and regenerated interphases in five successive rounds of phase partitioning, rather

than three.

Culture preparation of human CD4+ T cell blasts 19 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Blood (120 ml) was collected by venepuncture from two times four donors for RNA-IC and

OOPS. Peripheral blood mononuclear cells (PBMCs) were isolated by standard Ficoll

gradient (PancollR) centrifugation and CD4+ T cells were isolated from 1x108 cells using

CD4+ Microbeads (Miltenyi) to arrive at 2-3x107 cells. These were resuspended at 2x106/ml

in T cell medium ((AIM-V (Invitrogen), 10% heat-inactivated human serum, 2mM L-

glutamine, 10 mM HEPES and 1,25 µg/ml Fungizone), supplemented with 500 ng/ml PHA

(Murex) and 100 IU IL-2/ml (Novartis). Cells were dispensed in 24-well plates at 2 ml per

well and on day 3 old medium was replaced by fresh T cell medium and cell were

harvested and counted at day 4,5. For generating iTreg, naïve CD4+ T cells were isolated

from PBMCs using Microbeads (Naive CD4+ T Cell Isolation Kit II, Miltenyi) and

resuspended at 1x106 cells/ml in T cell medium supplemented with 500 ng/ml PHA (Murex),

500 IU IL-2/ml (Novartis), 500 nM Rapamycin (Selleckchem), and 5 ng/ml TGF-ß1

(Miltenyi). Cells were cultured in 24-well plates at 2 ml per well together with 1x106

irradiated (40 Gy) PBMCs derived from three donors. On day 3, old medium was replaced

by fresh T cell medium including supplements and the cells harvested and counted at day

5.

SDS-PAGE, Western blotting and silver staining

All protein visualization procedures were performed according to standard protocols. For

silver staining we used the SilverQuest kit (LC6070, Invitrogen). Antibodies used were anti-

GFP, (1:10, clone: 3E5-111, in house), anti-Ptbp1, (1:1000, 8776, Cell Signaling), anti-β-

tubulin, (1:1000, 86298, Cell Signaling). Proteins were visualized by staining with anti-rat

(1:3000, 7077, Cell Signaling) or anti-mouse (1:3000, 7076, Cell Signaling) secondary

antibodies conjugated to HRP.

Sample preparation for mass spectrometry

For RNA-capture, eluates were incubated with 10 µg/ml RNase A in 100 mM Tris, 50 mM

NaCl, 1 mM EDTA at 37°C for 30 min. RNase-treated eluates were acetone precipitated and

20 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

resuspended in denaturation buffer (6 M urea, 2 M thiourea, 10 mM Hepes, pH 8), reduced

with 1 mM DTT and alkylated with 5.5 mM IAA. Samples were diluted 1:5 with 62.5 mM Tris,

pH 8.1 and proteins digested with 0.5 µg Lys-C and 0.5 µg Trypsin at room temperature

overnight. The resulting peptides were desalted using stage-tips containing three layers of

C18 material (Empore).

For OOPS experiments, 100 µl of lysis buffer (Preomics, iST kit) were added and samples

incubated at 100°C for 10 min at 1,400 rpm. Samples were sonicated for 15 cycles (30s

on/30s off) on a bioruptor (Diagenode). Protein concentration was determined using the BCA

assay and about 30 µg of proteins were digested. To this end, trypsin and Lys-C were added

in a 1:100 ratio, samples diluted with lysis buffer to contain at least 50 µl of volume and

incubated overnight at 37°C. To 50 µl of sample, 250 ul Isopropanol/1% TFA were added

and samples vortexed for 15s. Samples were transferred on SDB-RPS (Empore) stagetips

(3 layers), washed twice with 100 µl Isopropanol/1% TFA and twice with 100 µl 0.2% TFA.

Peptides were eluted with 80 µl of 2% ammonia/80% acetonitrile, evaporated on a

centrifugal evaporator and resuspended with 10 µl of buffer A* (2% ACN, 0.1% TFA).

LC-MS/MS analysis

Peptides were separated on a reverse phase column (50 cm length, 75 μm inner diameter)

packed in-house with ReproSil-Pur C18-AQ 1.9 µm resin (Dr. Maisch GmbH). Reverse-

phase chromatography was performed with an EASY-nLC 1000 ultra-high pressure system,

coupled to a Q-Exactive HF Mass Spectrometer (Thermo Scientific) for mouse RNA-capture

experiments or a Q-Exactive HF-X Mass Spectrometer (Thermo Scientific) for human RNA-

capture, OOPS experiments and single-shot proteomes. Peptides were loaded with buffer A

(0.1% (v/v) formic acid) and eluted with a nonlinear 120-min (100-min gradient for human

RNA-capture and OOPS experiments) gradient of 5–60% buffer B (0.1% (v/v) formic acid,

80% (v/v) acetonitrile) at a flow rate of 250 nl/min (300 nl/min for human RNA-capture and

OOPS). After each gradient, the column was washed with 95% buffer B and re-equilibrated

with buffer A. Column temperature was kept at 60° C by an in-house designed oven with a

21 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Peltier element and operational parameters were monitored in real time by the SprayQc

software. MS data were acquired using a data-dependent top 15 (top 12 for human RNA-

capture and OOPS experiments) method in positive mode. Target value for the full scan MS

spectra was 3 × 106 charges in the 300−1,650 m/z range with a maximum injection time of

20 ms and a resolution of 60,000. The precursor isolation windows was set to 1.4 m/z and

capillary temperature was 250°C. Precursors were fragmented by higher-energy collisional

dissociation (HCD) with a normalized collision energy (NCE) of 27. MS/MS scans were

acquired at a resolution of 15,000 with an ion target value of 1 × 105, a maximum injection

time of 120 ms (60 ms for human RNA-capture and OOPS experiments). Repeated

sequencing of peptides was minimized by a dynamic exclusion time of 20 s (30 ms for

human RNA-capture and OOPS).

Raw data processing

MS raw files were analyzed by the MaxQuant software50 (version 1.5.1.6 for RNA-capture

files and version 1.5.6.7 for OOPS files) and peak lists were searched against the mouse or

human Uniprot FASTA database, respectively, and a common contaminants database (247

entries) by the Andromeda search engine51. Cysteine carbamidomethylation was set as fixed

modification, methionine oxidation and N-terminal protein acetylation as variable

modifications. False discovery rate was 1% for both proteins and peptides (minimum length

of 7 amino acids). The maximum number of missed cleavages allowed was 2. Maximal

allowed precursor mass deviation for peptide identification was 4.5 ppm after time-

dependent mass calibration and maximal fragment mass deviation was 20 ppm. Relative

quantification was performed using the MaxLFQ algorithm52. “Match between runs” was

activated with a retention time alignment window of 20 min and a match time window of 0.5

min for RNA-capture experiments, while matching between runs was disabled for OOPS

experiments. The minimum ratio count was set to 2 for label-free quantification.

22 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Data analysis

Statistical analysis of MS data was performed using Perseus (version 1.6.0.28). Human

RNA-capture, mouse RNA-capture, human OOPS and mouse OOPS data was processed

separately. For all experiments, MaxQuant output tables were filtered to remove protein

groups matching the reverse database, contaminants or proteins only observed with

modified peptides. Next, protein groups were filtered to have at least two valid values in

either the crosslinked or control triplicate. LFQ intensities were logarithmized (base 2) and

missing values were imputed from a normal distribution with a downshift of 1.8 standard

deviations and a width of 0.2 (0.25 for OOPS data). For RNA-capture experiments, a

Student’s T-test was performed to find proteins significantly enriched in the crosslinked

sample over the non-crosslinked control (false-discovery rate (FDR) < 0.05). As many

proteins were identified specifically in the crosslinked sample at intensities too low to find

significant differences compared to the imputed values, we additionally considered proteins

only identified in two or three replicates of the crosslinked sample, but never in the non-

crosslinked control, as RNA-binding proteins. For OOPS experiments, proteins significantly

enriched in a Student’s T-tests of the organic phase after RNase-treatment over the same

sample of the non-crosslinked control (FDR < 0.05) were considered RBPs.

GO term enrichment analysis was performed using the clusterProfiler package in R (version

4.1.0) as described in the original publication53. The mouse or human proteomes served as

background for the respective enrichment analysis. Relative abundance of proteins in the

single-shot proteome (Fig. 4a-b) was determined by calculating the logarithm (base 2) of the

ratio of the LFQ intensity and the number of theoretical peptides. Relative abundance of

proteins significant in the mouse OOPS and RNA-IC is shown in the single-shot proteome of

mouse CD4+ T cells measured with the OOPS samples. Relative abundance of proteins

significant in the human OOPS and RNA-IC experiments is shown in the single-shot

proteome of human CD4+ T cells measured with the OOPS samples or RNA-IC samples,

respectively. For the 4-way Venn comparison to find RBPs identified in more than one RNA-

IC or OOPS experiment, human gene names were converted into their homologous mouse

23 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

counterparts. Protein group entries containing more than one isoform were expanded for the

comparison and subsequently collapsed into one entry again for calculation of the Venn

diagram. Multiple gene name entries for different unambiguously identified isoforms were

collapsed into the major isoform. This affected six protein groups in the human RNA-IC

dataset and three protein groups in the mouse RNA-IC dataset.

Intrinsically disordered regions were retrieved from the Disorder Atlas54. We referred to the

LCR-eXXXplorer55 to obtain low complexity regions in proteins.

Plasmid construction

To generate a vector that expresses N-terminally GFP-tagged proteins, we amplified the

respective genes from cDNA of Teff or TFoxp3+ cells, added HindIII and KpnI restriction sites in

front of the start codon and cloned them into the pCR™8/GW/TOPO® vector. We then used

HindIII and KpnI to insert a GFP sequence where we removed the bases for the stop codon.

The respective sequences were subsequently transferred to the expression vector pMSCV

via the gateway cloning technology. Only GFP-Roquin-1 (a kind gift of Vigo Heissmeyer)

was expressed from the vector pDEST14. For oligonucleotide sequences see

Supplementary Table 3.

Validation of RNA binding ability

HEK293T cells were transfected by calcium phosphate transfection with plasmids

expressing the respective proteins with an N-terminal GFP-tag or GFP alone. After three

days, cells were washed with PBS on plates, UV crosslinked (CL) as before or directly

scraped from the plates (nCL). Cell lysates were generated by flash-freezing pellets in liquid

nitrogen and incubation in NP-40 lysis buffer (150 mM NaCl, 1% NP-40, 50 mM Tris-HCl, pH

7.4, 5 mM EDTA, 1 mM DTT, 1 mM PMSF and protease inhibitor mixture (Complete,

Roche)). After lysis, extracts were cleared by centrifugation at 17000 g for 15 min at 4° C.

We then determined protein concentration via the BCA method and used 2-10 mg of protein

24 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

for the subsequent GFP immunoprecipitation, depending on transfection efficiency and

expected RNA-binding capacity. We pre-coupled 200 µl Protein-G beads (10004D,

Dynabeads Protein G, Invitrogen) with 20 µg antibody (anti-GFP, clone: 3E5-111, in house)

in PBS (1 h, RT), washed beads in lysis buffer, added them to cell lysates and incubated

with rotation for 4 h at 4 °C. Beads were then washed three times with IP wash buffer (50

mM Tris-HCl, pH 7.5) with decreasing salt (500 mM, 350 mM, 150 mM, 50 mM NaCl) and

SDS (0.05%, 0.035%, 0.015%, 0.005%) concentrations. Proteins and crosslinked RNAs

were eluted with 50 mM glycine, pH 2.2 at 70° C for 5 min. Lämmli buffer (4x) was added

and samples were divided for mRNA and protein detection and separated via SDS gel

electrophoresis (6% SDS gels for detection of mRNA samples and 9% gels to verify

immunoprecipitation efficiency). For RNA detection we blotted onto Nitrocellulose

membranes and for protein detection on PVDF membranes. After transfer, the Nitrocellulose

membrane was prehybridized with Church buffer (0.36 M Na2HPO4, 0.14 M NaH2PO4, 1 mM

EDTA, 7% SDS) for 30 min and then incubated for 4 h with Church buffer containing 40 nM

3´-and 5´-Biotin labeled oligo(dT)20 probe to anneal to the poly-A tail of the bound mRNA.

The membrane was washed twice with 1 x SSC, 0.5% SDS and twice with 0.5 x SSC, 0.5%

SDS. Bound mRNA was detected with the Chemiluminescent Nucleic Acid Detection Kit

Module (89880, Thermo Fisher) according to the manufacturer´s instructions.

Real-Time PCR

Total RNA of input was purified from lysates with Agencourt RNAClean XP Beads (A63987,

Beckman Coulter) according to the manufacturer´s instructions and eluted in nuclease-free

H2O. cDNA was synthesized from total input RNA and oligo(dT)-isolated RNA with the

QuantiTect Reverse Transcription Kit (205311, Qiagen). All qRT-PCRs were performed with

the SYBR green method. For Primer sequences see Supplementary Table 3.

RNA isolation, reverse transcription and quantitative RT PCR as shown in Figure 8d was

performed as published 56 using the universal probes systems (Roche). Primers for Rc3h1

(F: gagacagcaccttaccagca; R: gacaaagcgggacacacat; probe 22) and Hprt (F:

25 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Tgatagatccattcctatgactgtaga; R: aagacattctttccagttaaagttgag; probe 95) were efficiency

tested (both E=1.99).

Tethering assay

Hela cells were seeded in 24-well plates using 5×104 cells per well. Transfection was

performed the following day using Lipofectamine2000 (Invitrogen) and 300 ng of total

constructs. Each transfection consisted of 75 ng of luciferase reporter plasmid psiCHECK2

(Promega) or luciferase-5boxB plasmid psiCHECK2 -5boxB, 225 ng of pDEST12.2-λN fused

constructs. After 24 h, cells were harvested for luciferase activity assays using a Dual-

Luciferase Reporter Assay System (Promega). Renilla luciferase activity was normalized to

Firefly luciferase activity in each well to control for variation in transfection efficiency.

psiCHECK2 lacking boxB sites served as a negative control, and each transfection was

analysed in triplicates.

BioID

The proximity-dependent biotin identification assay was performed according to Roux (Roux

et al., 2012) with modifications. For each sample 2x107 MEF cells were grown on ten 15-cm

cell culture dishes for 24h before BirA*-Roquin-1 or BirA* expression was induced by

doxycycline treatment. For T cells, transduction with the same BirA*-fusions cloned into the

plasmid pRetroXtight was performed as described above and the same number of cells was

used for the experiment. Six hours after addition of doxyclycline, biotin was added for 16h to

arrive at an end concentration of 50 µM. Approximately 8x107 cells per sample were

trypsinized, washed twice with PBS and lysed in 5 ml lysis buffer (50 mM Tris-HCL, ph7,4;

500 mM NaCl, 0,2% SDS; 1x protease inhibitors (Roche), 20 mM DTT, 25 U/ml Benzonase)

for 30 min at 4 °C using an end-over-end mixer. After adding 500 µl of 20% Triton X-100 the

samples were sonicated for two sessions of 30 pulses at 30% duty cycle and output level 2,

using a Branson Sonifier 450 device. Keep on ice for two minutes in between sessions.

26 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Pipetting of 4,5 ml prechilled 50 mM Tris-HCL, pH 7,4 was followed by an additional round of

sonication. During centrifugation at 16500 x g for 10 min at 4 °C 500 µl streptavidin beads

(Invitrogen) or each sample was equilibrated in a 1:1 mixture of lysis buffer and 50 mM Tris-

HCL pH 7,4. After overnight binding on a rotator at 4 °C Streptavidin beads were stringently

washed using wash buffers 1, 2 and 3 (Roux et al, 2012) and prepared for mass

spectrometry by three additional washes with buffer 4 (1 mM EDTA, 20 mM NaCl, 50 mM

Tris-HCL pH 7,4). Proteins were eluted from streptavidin beads with 50 µl of biotin-saturated

1x sample buffer (50 mM Tris-HCL pH 6,8, 12% sucrose, 2% SDS, 20 mM DTT, 0,004%

Bromphenol blue, 3 mM Biotin) by incubation for 7 min at 98 °C. For identification and

quantification of proteins, samples were proteolysed by a modified filter aided sample

preparation 57 and eluted peptides were analysed by LC-MSMS on a QExactive HF mass

spectrometer (ThermoFisher Scientific) coupled directly to a UItimate 3000 RSLC nano-

HPLC (Dionex). Label-free quantification was based on peptide intensities from extracted ion

chromatograms and performed with the Progenesis QI software (Nonlinear Dynamics,

Waters). Raw files were imported and after alignment, filtering and normalization, all MSMS

spectra were exported and searched against the Swissprot mouse database (16772

sequences, Release 2016_02) using the Mascot search engine with 10 ppm peptide mass

tolerance and 0.02 Da fragment mass tolerance, one missed cleavage allowed, and

carbamidomethylation set as fixed modification, methionine oxidation and asparagine or

glutamine deamidation allowed as variable modifications. A Mascot-integrated decoy

database search calculated an average false discovery of < 1% when searches were

performed with a mascot percolator score cut-off of 13 and significance threshold of 0.05.

Peptide assignments were re-imported into the Progenesis QI software. For quantification,

only unique peptides of an identified protein were included, and the total cumulative

normalized abundance was calculated by summing the abundances of all peptides allocated

to the respective protein. A t-test implemented in the Progenesis QI software comparing the

normalized abundances of the individual proteins between groups was calculated and

corrected for multiple testing resulting in q values (FDR adjusted p values) given in

27 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Supplemental Table 2. Only proteins identified by two or more peptides were included in

the list of Roquin-1 proximal proteins.

Antibodies

To generate monoclonal antibodies against pan-Ythdf proteins or λN-peptide, Wistar rats

were immunized with purified GST-tagged full-length mouse Ythdf3 protein or ovalbumin-

coupled peptide against λN (MNARTRRRERRAEKQWKAAN) using standard procedures as

described 58. The hybridoma cells of Ythdf- or λN-reactive supernatants were cloned at least

twice by limiting dilution. Experiments in this study were performed with anti-pan-Ythdf clone

DF3 17F2 (rat IgG2a/κ) and anti-λN clone LAN 4F10 (rat IgG2b/κ).

References

1. Hoefig, K.P. & Heissmeyer, V. Degradation of oligouridylated histone mRNAs: see UUUUU and goodbye. Wiley interdisciplinary reviews. RNA 5, 577-589 (2014).

2. Hoefig, K.P. & Heissmeyer, V. Posttranscriptional regulation of T helper cell fate decisions. The Journal of cell biology (2018).

3. Salerno, F., Turner, M. & Wolkers, M.C. Dynamic Post-Transcriptional Events Governing CD8(+) T Cell Homeostasis and Effector Function. Trends in immunology 41, 240-254 (2020).

4. Shulman, Z. & Stern-Ginossar, N. The RNA modification N(6)-methyladenosine as a novel regulator of the immune system. Nature immunology 21, 501-512 (2020).

5. Turner, M. & Hodson, D.J. An emerging role of RNA-binding proteins as multifunctional regulators of lymphocyte development and function. Adv Immunol 115, 161-185 (2012).

6. Chen, J. et al. Posttranscriptional gene regulation of IL-17 by the RNA-binding protein HuR is required for initiation of experimental autoimmune encephalomyelitis. Journal of immunology 191, 5441-5450 (2013).

7. Moore, M.J. et al. ZFP36 RNA-binding proteins restrain T cell activation and anti-viral immunity. eLife 7 (2018).

28 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

8. Stellato, C. et al. Coordinate regulation of GATA-3 and Th2 cytokine by the RNA-binding protein HuR. Journal of immunology 187, 441-449 (2011).

9. Vogel, K.U., Bell, L.S., Galloway, A., Ahlfors, H. & Turner, M. The RNA-Binding Proteins Zfp36l1 and Zfp36l2 Enforce the Thymic beta-Selection Checkpoint by Limiting DNA Damage Response Signaling and Cell Cycle Progression. Journal of immunology 197, 2673-2685 (2016).

10. Jeltsch, K.M. et al. Cleavage of roquin and regnase-1 by the paracaspase MALT1 releases their cooperatively repressed targets to promote T(H)17 differentiation. Nature immunology 15, 1079-1089 (2014).

11. Vogel, K.U. et al. Roquin paralogs 1 and 2 redundantly repress the Icos and Ox40 costimulator mRNAs and control follicular helper T cell differentiation. Immunity 38, 655-668 (2013).

12. Minagawa, K. et al. Posttranscriptional modulation of cytokine production in T cells for the regulation of excessive inflammation by TFL. Journal of immunology 192, 1512-1524 (2014).

13. Uehata, T. et al. Malt1-induced cleavage of regnase-1 in CD4(+) helper T cells regulates immune activation. Cell 153, 1036-1049 (2013).

14. Baumjohann, D. & Ansel, K.M. MicroRNA-mediated regulation of T helper cell differentiation and plasticity. Nature reviews. Immunology 13, 666-678 (2013).

15. Akira, S. Regnase-1, a ribonuclease involved in the regulation of immune responses. Cold Spring Harb Symp Quant Biol 78, 51-60 (2013).

16. Heissmeyer, V. & Vogel, K.U. Molecular control of Tfh-cell differentiation by Roquin family proteins. Immunological reviews 253, 273-289 (2013).

17. Hodson, D.J., Screen, M. & Turner, M. RNA-binding proteins in hematopoiesis and hematological malignancy. Blood 133, 2365-2373 (2019).

18. Turner, M., Galloway, A. & Vigorito, E. Noncoding RNA and its associated proteins as regulatory elements of the immune system. Nature immunology 15, 484-491 (2014).

19. Rehage, N. et al. Binding of NUFIP2 to Roquin promotes recognition and regulation of ICOS mRNA. Nature communications 9, 299 (2018).

20. Mukherjee, N. et al. Global target mRNA specification and regulation by the RNA- binding protein ZFP36. Genome biology 15, R12 (2014).

21. Masuda, K. et al. Arid5a controls IL-6 mRNA stability, which contributes to elevation of IL-6 level in vivo. Proceedings of the National Academy of Sciences of the United States of America 110, 9409-9414 (2013).

29 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

22. Gerstberger, S., Hafner, M. & Tuschl, T. A census of human RNA-binding proteins. Nature reviews. Genetics 15, 829-845 (2014).

23. Castello, A. et al. Insights into RNA biology from an atlas of mammalian mRNA- binding proteins. Cell 149, 1393-1406 (2012).

24. Hentze, M.W., Castello, A., Schwarzl, T. & Preiss, T. A brave new world of RNA- binding proteins. Nat Rev Mol Cell Biol 19, 327-341 (2018).

25. Liao, J.Y. et al. EuRBPDB: a comprehensive resource for annotation, functional and oncological investigation of eukaryotic RNA binding proteins (RBPs). Nucleic acids research 48, D307-D313 (2020).

26. Glasmacher, E. et al. Roquin binds inducible costimulator mRNA and effectors of mRNA decay to induce microRNA-independent post-transcriptional repression. Nature immunology 11, 725-733 (2010).

27. Pratama, A. et al. MicroRNA-146a regulates ICOS-ICOSL signalling to limit accumulation of T follicular helper cells and germinal centres. Nature communications 6, 6436 (2015).

28. Vinuesa, C.G. et al. A RING-type ubiquitin ligase family member required to repress follicular helper T cells and autoimmunity. Nature 435, 452-458 (2005).

29. Zhu, Y. et al. The E3 ligase VHL promotes follicular helper T cell differentiation via glycolytic-epigenetic control. The Journal of experimental medicine 216, 1664-1681 (2019).

30. Zaccara, S., Ries, R.J. & Jaffrey, S.R. Reading, writing and erasing mRNA methylation. Nat Rev Mol Cell Biol 20, 608-624 (2019).

31. Edupuganti, R.R. et al. N(6)-methyladenosine (m(6)A) recruits and repels proteins to regulate mRNA homeostasis. Nature structural & molecular biology 24, 870-878 (2017).

32. Liu, N. et al. N(6)-methyladenosine-dependent RNA structural switches regulate RNA-protein interactions. Nature 518, 560-564 (2015).

33. Weichmann, F. et al. Validation strategies for antibodies targeting modified ribonucleotides. RNA (2020).

34. Rao, P.K. et al. Loss of cardiac microRNA-mediated regulation leads to dilated cardiomyopathy and heart failure. Circ Res 105, 585-594 (2009).

35. Sledzinska, A. et al. TGF-beta signalling is required for CD4(+) T cell homeostasis but dispensable for regulatory T cell function. PLoS Biol 11, e1001674 (2013).

30 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

36. Bronevetsky, Y. et al. T cell activation induces proteasomal degradation of Argonaute and rapid remodeling of the microRNA repertoire. The Journal of experimental medicine 210, 417-432 (2013).

37. Bohjanen, P.R., Moua, M.L., Guo, L., Taye, A. & Vlasova-St Louis, I.A. Altered CELF1 binding to target transcripts in malignant T cells. RNA 21, 1757-1769 (2015).

38. Castello, A. et al. Comprehensive Identification of RNA-Binding Domains in Human Cells. Mol Cell 63, 696-710 (2016).

39. Queiroz, R.M.L. et al. Comprehensive identification of RNA-protein interactions in any organism using orthogonal organic phase separation (OOPS). Nat Biotechnol 37, 169-178 (2019).

40. Villanueva, E. et al. Efficient recovery of the RNA-bound proteome and protein- bound transcriptome using phase separation (OOPS). Nat Protoc (2020).

41. Amaya Ramirez, C.C., Hubbe, P., Mandel, N. & Bethune, J. 4EHP-independent repression of endogenous mRNAs by the RNA-binding protein GIGYF2. Nucleic acids research 46, 5792-5808 (2018).

42. Roux, K.J., Kim, D.I., Raida, M. & Burke, B. A promiscuous biotin ligase fusion protein identifies proximal and interacting proteins in mammalian cells. The Journal of cell biology 196, 801-810 (2012).

43. Leppek, K. et al. Roquin promotes constitutive mRNA decay via a conserved class of stem-loop recognition motifs. Cell 153, 869-881 (2013).

44. Sgromo, A. et al. A CAF40-binding motif facilitates recruitment of the CCR4-NOT complex to mRNAs targeted by Drosophila Roquin. Nature communications 8, 14307 (2017).

45. Peyman, J.A. Repression of major histocompatibility complex genes by a human trophoblast ribonucleic acid. Biol Reprod 60, 23-31 (1999).

46. Peyman, J.A. Mammalian expression cloning of two human trophoblast suppressors of major histocompatibility complex genes. Am J Reprod Immunol 45, 382-392 (2001).

47. Keene, J.D. RNA regulons: coordination of post-transcriptional events. Nature reviews. Genetics 8, 533-543 (2007).

48. Tavernier, S.J. et al. A human immune dysregulation syndrome characterized by severe hyperinflammation with a homozygous nonsense Roquin-1 mutation. Nature communications 10, 4779 (2019).

31 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

49. Li, Y. et al. Central role of myeloid MCPIP1 in protecting against LPS-induced inflammation and lung injury. Signal Transduct Target Ther 2, 17066 (2017).

50. Cox, J. & Mann, M. MaxQuant enables high peptide identification rates, individualized p.p.b.-range mass accuracies and proteome-wide protein quantification. Nat Biotechnol 26, 1367-1372 (2008).

51. Cox, J. et al. Andromeda: a peptide search engine integrated into the MaxQuant environment. J Proteome Res 10, 1794-1805 (2011).

52. Cox, J. et al. Accurate proteome-wide label-free quantification by delayed normalization and maximal peptide ratio extraction, termed MaxLFQ. Mol Cell Proteomics 13, 2513-2526 (2014).

53. Yu, G., Wang, L.G., Han, Y. & He, Q.Y. clusterProfiler: an R package for comparing biological themes among gene clusters. OMICS 16, 284-287 (2012).

54. Vincent, M. & Schnell, S. Disorder Atlas: Web-based software for the proteome- based interpretation of intrinsic disorder predictions. Comput Biol Chem 83, 107090 (2019).

55. Kirmitzoglou, I. & Promponas, V.J. LCR-eXXXplorer: a web platform to search, visualize and share data for low complexity regions in protein sequences. Bioinformatics 31, 2208-2210 (2015).

56. Hoefig, K.P. et al. Eri1 degrades the stem-loop of oligouridylated histone mRNAs to induce replication-dependent decay. Nature structural & molecular biology 20, 73-81 (2013).

57. Grosche, A. et al. The Proteome of Native Adult Muller Glial Cells From Murine Retina. Mol Cell Proteomics 15, 462-480 (2016).

58. Feederle, R. et al. Generation of Pax1/PAX1-Specific Monoclonal Antibodies. Monoclon Antib Immunodiagn Immunother (2016).

59. Sysoev, V.O. et al. Global changes of the RNA-bound proteome during the maternal- to-zygotic transition in Drosophila. Nat Commun 7, 12128 (2016).

Figure legends

Figure 1: Icos responds to inputs from several post-transcriptional regulators. a, d, g,

j, Bar diagrams of Icos mean fluorescence intensities (MFI) over a five-day period of T cell

32 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

activation illustrating changes induced by the CD4+-specific, induced depletion of the

respective genes. Significance was calculated using the unpaired t-test (two-tailed) with n=3.

b, e, h, k, Representative histograms of Icos expression at the specified days. c, f, i, l,

Histograms demonstrating successful depletion of the respective target proteins. o-r,

Western blots showing patterns of dynamic RBP regulation after CD3 and CD28 CD4+ T cell

activation. ns = not significant; * significant (p value = 0.01 – 0.05); ** very significant (p

value = 0.001 – 0.01); *** very significant (p value = < 0.001).

Figure 2: The CD4+ T cell RBPome of polyadenylated RNAs. a, Schematic illustration of

the RNA-interactome capture (RNA-IC) method that was carried out to identify RBPs from

mouse and human CD4+ T cells. b, c, Volcano plot showing the -log10 p-value plotted

against the log2 fold-change comparing the RNA-capture from crosslinked (CL) mouse CD4

T cells (b) or human CD4+ T cells (c) versus the non-crosslinked (nCL) control. Red dots

represent proteins significant at a 5% FDR cut-off level in both mouse and human RNA-

capture experiments and blue dot proteins were significant only in mouse or human,

respectively. d, Enrichment analysis of GO Molecular Function terms of significant proteins

in mouse or human RNA capture data. The 10 most enriched terms in mouse (dark blue)

and the respective terms in human (light blue) are shown. The y-axis represents the number

of proteins matching the respective GO term. Numbers above each term depict the adjusted

p-value after Benjamini-Hochberg multiple testing correction. e, Distribution of IDRs in all

Uniprot reviewed protein sequences (black line), in proteins of the mouse EuRBPDB

database (green line) and in proteins significant in the mouse RNA-IC experiment (red line).

The same plot is shown for human data at the bottom. According to Kolmogorov-Smirnov

testing the IDR distribution differences between RNA-IC (red lines) and all proteins (black

lines) are highly significant in mouse and man and reach the smallest possible p-value

(p<2.2x10-16).

Figure 3: The global RNA-bound proteome of CD4+ T cells. a, Schematic overview of the

OOPS method 39 with phase partitioning cycles increased to five. b-c, Volcano plots showing

33 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

the -log10 p-value plotted against the log2 fold-change comparing the organic phase after

RNase treatment of the interphase of OOPS experiments of crosslinked mouse CD4+ T cells

(b) or human CD4+ T cells (c) versus the same non-crosslinked sample. Red dots represent

proteins significant at a 5% FDR cutoff level in both mouse and human OOPS experiments

and blue dots proteins significant only in mouse or human, respectively. d, Enrichment

analysis of GO Molecular Function terms of significant proteins in mouse or human OOPS

data. Enriched terms are depicted exactly as described for RNA-capture data in Figure 2d. e,

Distribution of intrinsically disordered regions in all Uniprot reviewed protein sequences

(black line), in proteins in the mouse EuRBPDB database (green) and in proteins significant

in the mouse OOPS data (red line). The same plot is shown for human data at the bottom.

According to Kolmogorov-Smirnov testing the IDR distribution differences between RNA-IC

(red lines) and all proteins (black lines) are highly significant in mouse and man and reach

the smallest possible p-value (p<2.2x10-16).

Figure 4: Defining the mouse and human CD4+ T cell RBPomes.

a, Relative abundance (Log2) of proteins identified in a single-shot mouse proteome (orange

dots) plotted by their rank from highest to lowest abundant protein. RNA-binding proteins

detected by RNA capture (top plots) or by OOPS (bottom plots) are highlighted as blue

diamonds. b, Same plots as shown in (a) for human data. c, e, The recently established

database EuRBPDB was used as a reference for eukaryotic RBPs to determine the numbers

of canonical and non-canonical RBPs identified by RNA-IC or OOPS on mouse (c) or human

(e) CD4+ T cells. d, f, The occurrence of RBDs in RNA-IC and OOPS-identified RBPs was

analyzed in comparison with the ten most abundant motifs described for the mouse (d) or

human (f) proteome. g, Venn diagram using four datasets for RBPs in CD4+ T cells as

determined by RNA-IC and OOPS in mouse and human cells. *The core RBPomes contain

proteins present in at least two datasets.

Figure 5: Stat1 and Stat4 are RBPs with regulatory potential. a, b, Semi-quantitative

identification method for RBPs as in 59. In short, GFP-fused proteins were transfected into

34 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

HEK293T cells, crosslinked or left untreated and subsequently immunoprecipitated using an

anti-GFP antibody. The obtained samples were divided for protein and RNA detection by

Western or Northern blotting, respectively. c, Schematic representation of the tethering

assay that was used to investigate a possible influence of the genes of interest (GOI) on

renilla luciferase expression. The affinity of the λN peptide targets the respective fusion

protein to boxB stem-loop structures (5x) in the 3’-UTR of a Renilla luciferase gene where it

can exert its function, if exciting. d, FACS plots using the new monoclonal antibody 4F10

demonstrate specific λN detection of a λN-GFP fusion protein expressed in 293T cells. e,

Western blot showing the expression of the indicated λN-GOI proteins in 293T cells after

transfection. f, Tethering assay results performed in HeLa cells as explained in (c). Two

negative controls were implemented, using constructs without boxB sites or λN expression

without fusion to a GOI. Each measurement was performed in triplicate and was

independently repeated at least twice (n=2).

Figure 6: Identification of proteins in proximity to Roquin-1 in CD4+ T cells. a,

Schematic overview of the BioID method showing how addition of biotin to the medium leads

to the activation of biotin, diffusion of biotinoyl-5’AMP and the biotinylation of the bait

(Roquin-1) and all preys in the circumference. b, Equimolar amounts of protein were loaded

onto a PAGE gel for Western blotting applying an anti-biotin antibody. Efficient biotinylation

of both baits BirA*-Roquin-1 and BirA*-GFP (control) could be demonstrated. c, Histogram

showing that transduction of CD4+ T cells with retrovirus to inducibly express BirA*-Roquin-1

lead to the efficient downregulation of endogenous Icos. d, Identified preys from Roquin-1

BioIDs (n=5) in CD4+ T cells. Depicted are all significantly enriched proteins with the

exception of highly abundant ribosomal and histone proteins. Dot sizes equal p-values and

positioning towards the center implies increased x-fold enrichment over BirA*-GFP BioID

results. e, Venn diagram showing the overlap of RBPs from the CD4+ T cell RBPome with

the proteins identified by Roquin-1 BioID in T cells. f, Listed are all 38 proteins (59%) that

35 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

are Roquin-1 preys and RBPs. Colored fields indicate proteins that were chosen for further

analysis.

Figure 7: Higher order Icos regulation by Roquin-1 and Celf1. Treatment with 4’-OH-

tamoxifen of CD4+ T cell with the genotypes Rc3h1fl/fl;Rc3h2fl/fl;rtTA3 without (WT) or with the

CD4Cre-ERT2 allele (iDKO) were used for transduction with retroviruses. Expression levels

of Icos and four additional Roquin-1 targets in WT and iDKO cells were analyzed 16h after

individual, doxycycline-induced overexpression of 46 GFP-GOI fusion genes. a, c, e, g, The

geometrical mean of each GOI-GFP divided by GFP for each Roquin-1 target was calculated

and the summarized results are shown as bar diagrams for (a) Vav1, (b) Rbms1, (c) Cpeb4

and (d) Celf1. b, d, f, h, Representative, original FACS data are depicted as histograms or

contour plots. Experiments for the negative Vav1 result were repeated twice, those for

Rbms1, Cpeb4 and Celf1 at least three times.

Supplementary Fig. 1: RNA-IC supporting results.

a, Western blots demonstrating specificity of the newly established pan-Ythdf monoclonal

antibody 17F2 for N-terminal GFP fusions of Ythdf1, Ythdf2 and Ythdf3, which are absent

from non-transfected 293T cells. b, Silver staining analysis of oligo(dT) captured samples

with and without UV irradiation. c, Quantitative RT-PCR to determine RNA pull-down

efficiency with (CL) and without crosslink (nCL). Error bars show the standard deviation

around the means of three independent experiments. d, Western blotting of UV irradiated

and nonirradiated samples of EL-4 T cells. Membranes were probed with antibodies for the

known mRNA-binding protein polypyrimidine tract binding protein 1 (Ptbp1) and β-tubulin. e,

Distribution of low complexity regions in all Uniprot reviewed protein sequences (black line),

in proteins in the EuRBPDB database (green) and in proteins significant in the RNA-IC data

(red line). The left plot shows mouse data and the right plot human data. According to

Kolmogorov-Smirnov testing the LCR distribution differences between RNA-IC (red lines)

and all proteins (black lines) are highly significant in mouse (p<2.2x10-16) and man

(p=7.8x10-16).

36 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Supplementary Fig. 2: RNA-IC on mouse and human iTreg cells. a, Volcano plot

showing the -log10 p-value in relation to the log2 fold-change comparing the RNA-capture

from crosslinked mouse regulatory T cells (left plot) or human regulatory T cells (right plot)

versus the non-crosslinked control. Red dots represent proteins significant at a 5% FDR

cutoff level in both mouse and human RNA-capture experiments and blue dots proteins

significant only in mouse or human, respectively. b, Venn diagram using four datasets to

compare RNA-IC derived RBPomes of effector T and induced Treg cells from mouse and

man.

Supplementary Fig. 3: OOPS supporting results. a, Agarose gel demonstrating

disappearance of the typical 18S/28S rRNA bands after crosslinking and appearance of

upshifted protein-RNA adducts (black arrows) in MEF cells with and without doxycycline-

induce Roquin-1 expression. rRNA bands reappear after proteinase K (PK) treatment (grey

arrows) and after RNase treatment the protein-RNA adducts in the wells disappear. b,

Western blots showing that known RBPs, such as Roquin-1 and Gapdh can be detected in

interphases, in the case of Roquin-1 only after induced expression. * cleavage product. c, d,

Volcano plot showing the -log10 p-value plotted against the log2 fold-change comparing the

interphase of OOPS experiments of crosslinked mouse CD4+ T cells versus the organic

phase after RNase treatment of the interphase. Glycoproteins are highlighted in red.

Enrichment analysis of GO Molecular Function terms was performed for proteins with a log2

fold-change larger than 2. The 20 most enriched terms are depicted.

Supplementary Figure 4: Gene ontology enrichment analysis. Enrichment analysis of

GO Biological Process and GO Cellular Component terms of significant proteins in mouse or

human RNA-IC data (top row) or OOPS data (bottom row). The ten most enriched terms in

mouse (dark blue) and the respective terms in human (light blue) are shown. The y-axis

represents the number of proteins matching the respective GO term. Numbers above each

term depict the adjusted p-value after multiple testing correction (Benjamini-Hochberg).

37 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Supplementary Fig. 5: All currently annotated CCCH and KH domain containing RBPs

and their detection in the RNA-IC and OOPS-identified RBPomes. All RBPs with the

indicated RBDs are listed according to EuRBPDB. Colored boxes indicate significant

enrichment by RNA-IC or OOPS in mouse or human CD4+ T cells.

Supplementary Fig. 6: Identification of Roquin-1 preys by BioID in MEF cells. a,

Western blots showing doxycycline-induced expression of Myc-BirA*-Roquin-1 or Myc-BirA*

in MEF cell clones. b, FACS blot demonstrating that as in T cells (Fig. 6c) N-terminal fusion

of Myc-BirA* to Roquin-1 does not affect its function and c, the fusion protein can biotinylate

Roquin-1 and d, its cofactor Nufip2. e, Identified preys from Roquin-1 BioID in MEF cell

clones. Depicted are 55 of 143 significantly enriched proteins (n=4) for better comparison

with Fig. 6d. Yellow dot color indicates identification of the Roquin-1 prey in MEF and T cells.

Dot sizes equal p-values and positioning towards the center implies increased x-fold

enrichment.

Supplementary Fig. 7: Overlap between Roquin-1 preys in MEF cells and the CD4+ T

cell RBPome. a, Venn diagram showing that 96 proteins (67%) are Roquin-1 preys and also

RBPs in T cells. b, Table listing these 96 proteins. Colors indicate which genes were clone

for downstream experiments and if they were identified by MEF BioID only (red) or

additionally in the T cell BioID.(blue).

Supplementary Fig. 8: Supporting results for higher order Icos regulation by Roquin-1

and Celf1. a, Microscopical images showing different localizations of the GFP signal as a

result of GFP-GOI subcellular targeting in transfected 293T cells. b, Schematic

representation of the experiment performed in Fig. 7. c, Treatment with 4’-OH-tamoxifen of

CD4+ T cell with the genotypes Rc3h1fl/fl;Rc3h2fl/fl;rtTA3 without (WT) or with the CD4Cre-

ERT2 allele (iDKO) were used for transduction with a retrovirus expressing GFP-Roquin-1.

Expression levels of the Roquin-1 targets Icos, Ox40, Ctla4 and IκbNS were analyzed and

work as a positive control for the experiment. d, e, Cells taken from an experiment as

performed (b). qPCR (d) and Western blot (e) performed on cDNA or protein lysates,

38 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

respectively, derived from CD4+ T cells (WT) after doxycycline-induced expression of the

indicated fusion proteins.

39 a b c CD4-Cre-ERT2 (WT) Rc3h1-2f/f;CD4-Cre-ERT2 0d 4d 5 ** ns ** *** ** 100 100

) bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint

10 (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. 4 60 60

3 20 20 Icos (MFI log Cells (%) Cells (%) 2 3 3 4 5 3 3 4 5 3 3 4 5 0d 1d 2d 3d 4d 5d -10 0 10 10 10 -10 0 10 10 10 -10 0 10 10 10 Icos-FITC Roquin-AF647 d CD4-Cre-ERT2 (WT) e f Zc3h12f/f;CD4-Cre-ERT2 0d 4d 5 ns ns ns ns ns ns 100 100 ) 10 4 60 60

3 20 20 Icos (MFI log Cells (%) 2 Cells (%) -103 0 103 104 105 -103 0 103 104 105 -103 0 103 104 105 0d 1d 2d 3d 4d 5d Icos-FITC Regnase-1-AF647 g CD4-Cre-ERT2 (WT) h i Wtapf/f;CD4-Cre-ERT2 1d 5d 5 100 100 ns ns ns ns ns

) ** 10 4 60 60

3 20 20 Icos (MFI log Cells (%) 2 Cells (%) 3 3 4 5 3 3 4 5 3 3 4 5 0d 1d 2d 3d 4d 5d -10 0 10 10 10 -10 0 10 10 10 -10 0 10 10 10 Icos-FITC Wtap-FITC j CD4-Cre-ERT2 (WT) k l Dgcr8f/f;CD4-Cre-ERT2 1d 5d 5 100 ns ns ns ns * ** 100 ) 10 4 60 60

3 20 20 Icos (MFI log Cells (%) 2 Cells (%) 3 3 4 5 3 3 4 5 3 3 4 5 0d 1d 2d 3d 4d 5d -10 0 10 10 10 -10 0 10 10 10 -10 0 10 10 10 Icos-FITC Dgcr8-APC

m 0d 1d 2d 3d 4d 5d kDa n 0d 1d 2d 3d 4d 5d kDa 63 Roquin-2 TTP Roquin-1 135 100 Pan-Ythdf 75 75 Gapdh 35 579 510 63 o 0d 1d 2d 3d 4d 5d Regnase-1 75 Rbms1 48 Pan-Ago 100 Gapdh 35 0d 1d 2d 3d 4d 5d Nuf p2 75 p Cpeb4 75 Fmrp 75 Gapdh 35 Fxr1 75 q 0d 1d 2d 3d 4d 5d Celf1 Fxr2 75 48 Gapdh 35 Gapdh 35

Figure 1 a Differentiation UV crosslink Oligo(dT) Mass

into Teff (254 nm) capture Spec. bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowedAAAAA without permission. TTTTT Cell lysis Elution CD4 T cell Expansion + oligo(dT) RNase isolation beads digest

Poly(A)-mRNA

b c mouse RNA-IC human RNA-IC 7 8 SYNCRIP significant in mouse significant in human HNRNPC Pum2 7 6 and human G3bp1 Hnrnpm and mouse HNRNPL Hnrnpab Tial1 EIF3A significant in mouse Raly 6 significant in human HDLBP 5 Pum1 Ptbp3 DDX3X CSDE1 Rbms1 Mbnl1 5 RALY 4 Elavl1 ELAVL1 Celf1 4 CELF2 RC3H1 CELF1 3 Rc3h1 Cpeb4 RBMS1

3 p-value [-log10] p-value [-log10] 2 Vav1 Crip1 2 Ldha CPEB4 1 1

0 0

-10 -5.0 0 5.0 10 -10 -5.0 0 5.0 10 Student's T-Test Difference CL vs nCL [log2] Student's T-Test Difference CL vs nCL [log2]

d e IDR - RNA-IC 1.0 GO - Molecular function mRNA-IC mEuRBPDB 0.8 all proteins

80 2.1E-44 3.9E-40 0.6 70 mouse human 0.4 60

50 0.2

proportion of proteins

3.2E-18

4.4E-20

1.9E-23 1.1E-21 40 0.0

# of proteins

9.1E-11

4.1E-8

5.5E-10

4.1E-8

2.5E-10

3.9E-10

1.2E-7

6.7E-10

5.8E-8 9.1E-11

1.5E-8 3.9E-10 3.1E-9 30 5.0E-9 1.0 hRNA-IC hEuRBPDB 20 0.8 all proteins 10 0.6 0 0.4

0.2

mRNA binding miRNA binding proportion of proteins

poly(U) RNA binding 0.0 mRNA 3'-UTR binding regulatorypoly-purine RNA binding tract binding 0.0 0.2 0.4 0.6 0.8 1.0 poly-pyrimidine tract bindingtranslation regulator activity amount of region single-stranded RNA bindingdouble-stranded RNA binding

Figure 2 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. a b mouse 36 human 36 32 32 28 RNA-IC bioRxiv preprint28 doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights24 reserved. No reuse allowed without permission. [log2] 24 20 20

Rel. abundance 16 500 1500 2500 3500 500 1500 2500 3500 Rank Rank 36 36 32 32 OOPS 28 28

[log2] 24 24 20 20 Rel. abundance 500 1500 2500 3500 500 1500 2500 3500 Rank Rank = identified in single-shot proteome = significant in RNA-IC or OOPS

c d mEuRBPDB 4096 128 mOOPS 1024 64 mRNA-IC 256 32 64 16

16 8 # of RBPs #of # of RBPs of # 4 4 2

1 1

RIO1

KH_1

ResIII

zf-met DEAD

RRM_1 RRM_5 zf-C2H2

e f zf-CCCH MMR_HSR1 4096 hEuRBPDB 128 hOOPS 1024 64 hRNA-IC 256 32 64 16 8

16 # of RBPs of # # of RBPs of # 4 4 2

1 1

KH_1

ResIII

zf-met DEAD

RRM_1 RRM_5

zf-C2H2

zf-CCCH

MMR_HSR1 Methyltransf_31 RBD abundance in proteome g mOOPS hOOPS 1255 1159

mRNA-IC 519 424 hRNA-IC 312 308 11 486 14

35 23 27 24

8 161 8 16 20

Mouse RBPome: 1352 38 Human RBPome: 1250 *Core RBPome: 798 *Core RBPome: 801 Figure 4 a bioRxiv preprintUV doi:- https://doi.org/10.1101/2020.08.20.259234+- + - + ; this- version+ posted August 20, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. Roquin

GFP

Rbms1

Ldha

Protein mRNA b UV - + - + - + - + Stat1

Stat4

Crip1 0,2% Input 8% Capture 0,014% Input 92% Capture

Protein mRNA c d GFP λN-GFP 105 λN-GOI 104 ? 3 anti-λN

λ N 10 (4F10) Renilla Firefly poly(A) luciferase luciferase poly(A) 102 Promoter BoxB sites Promoter 101 101 102 103 104 105 GFP e f

no boxB 120 5 boxB

λN-Roquin-1(148λN-Vav-1(111λN-Crip1(22 kDa)λ N-Stat1(104kDa) λkDa)N-Stat4(101λ N-Celf1(65kDa)λ N-Rbms1(58kDa) kDa) λN-Cpeb4 kDa) λN-Pat1b (93 kDa) (108 kDa) kDa 135 80 100 RLU 75 anti-λN 63 (4F10) 40 48

35 0 25

anti-Gapdh

Figure 5 bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint

(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

(10x) (10x)

a b qi (10 oquin BirA*- (10 oquin

BirA-GFP BirA-GFPBirA-Roquin BirA-RoquinBirA-san BirA-san r-a (10x) irA-san

Dox-inducible BirA* fusion: Labelling radius BirA*-GFP Roquin-1 (10x) irA-san

(10x) (10x)

ID ID ut

(~ 10 nm) ut

oDBirA-GFP ioID oDBirA-GFP ioID

APd APd

APd APd

Input Input

Input Input

iI BirA-R BioID iI BirA-R BioID

Bio Bio

B B

Inp Inp

BioID BioID

BioID BioID

iI B BioID iI B BioID

APd APd

Input BioID FT Input BioID FT BioID BioID addition BirA*- of biotin 135 135 Roquin-1 anti-biotin BirA* Roquin-1 BirA* Roquin-1 100 100

75 75 BirA*-GFPn n

ti ti

o o

Bi Bi Roquin-1-binding proteins c Transduced T cells 100 Peripheral proteins BirA*- Biotinoyl-5‘AMP 80 Roquin-1 BirA*-GFP Covalently bound biotin 60 40 20

d Cells (%) Nufip2 Stat3 0 103 104 105 Ywhab Actb Ubap2 Icos-PE Sash3 Ubap2l Llph Helz Surf6 Pdcl3 Alyref Atxn2l Fam120a R3hdm2 Rc3h1 = bait Coro1b Pcca Pspc1 Cnot2 Prcc2b/a/c Cnot1 p < 0.001 Upf1 Plec Psmc2 Swap70 Rc3h1 p < 0.002 Rc3h2 Keap1 p < 0.004 Zap70 Tdrd3 100x Rrbp1 Ybx1 Ssb p < 0.02 Eif4enif1 Abcf1 Lsm14a 50x Ythdf3/1 p < 0.05 Rtn4 Mccc1 Pip4k2c 20x Tnrc6a/b Fnbp1 10x Pc Dmnt1 Gapdh Xrn1 5x Ebna1bp2 Ddx6 Edc4 Usp36 2x f Smg7 Roquin-1 T cell BioID overlap with RBPome

Expanded T cell Abcf1 Fnbp1 Prrc2c Tnrc6b e RBPome (1320) Alyref Helz Pspc1 Ubap2 Atxn2l Hist1h1c R3hdm2 Ubap2l Cnot1 Hist1h3b Rc3h1 Upf1 T cell Roquin-1 Cnot2 Lsm14a Rpl23a Xrn1 BioID (64) Ddx6 Nufip2 Smg7 Ybx1 1282 3826 Ebna1bp2 Pc Ssb Ythdf1 Edc4 Plec Stat3 Ythdf3 Eif4enif1 Prrc2a Tdrd3 Fam120a Prrc2b Tnrc6a cloned for induced expression

Figure 6 WT iDKO WT iDKO a Vav1 b Icos Icos 2,5bioRxiv preprint doi: https://doi.org/10.1101/2020.08.20.259234; this version posted August 20, 2020. The copyright holder for this preprint (which was notWT certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. 2,0 iDKO Icos-PE

1,5

100 5 1,0 Ctla4 Ctla4 10 80 4 10 0,5 NS 60 103 Icos Ctla4 Ox40 I κ B Cells (%) 40 2

Ctla4-PE 10 0,0 Mean Fluorescence (GOI/GFP) 20 1 Regnase-1 10 0 101102103104105 101102103104105 c Rbms1 d Icos Icos 2,5 WT 2,0 iDKO Icos-PE

1,5

1,0 100 Ctla4 Ctla4 105

80 104 NS 0,5 60 3

Icos 10 Ctla4 Ox40 I κ B

Cells (%) 40

Ctla4-PE 2 0,0 10 Mean Fluorescence (GOI/GFP) 20 Regnase-1 101 0 101102103104105 1 2 3 4 5 e Cpeb4 f 10 10 10 10 10 3,0 Ox40 Ox40 WT 2,5 iDKO

2,0 Ox40-APC

1,5 100 5 1,0 Ctla4 Ctla4 10 80 104 0,5

60 103 NS

Cells (%) 40 2 Ctla4-PE Mean Fluorescence (GOI/GFP) 0,0 Icos 10 Ctla4 Ox40 I κ B 20 101 0 5 Regnase-1 1 2 3 4 5 g Celf1 h 10 10 10 10 10 10110210310410 Icos Icos 2,5 WT 2,0 iDKO Icos-PE

1,5

100 5 1,0 Ctla4 Ctla4 10

80 104 0,5 NS 60 3

Icos 10 Ctla4 Ox40 I κ B

Cells (%) 40 2 0,0 Ctla4-PE 10 Mean Fluorescence (GOI/GFP) 20 1 Regnase-1 10 0 101102103104105 101102103104105 target expression GFP

WT+GFP-GOI iDKO+GFP-GOI T cell+GFP

Figure 7